This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

gg-mode of neutron stars in pseudo-Newtonian gravity

Hong-Bo Li Department of Astronomy, School of Physics, Peking University, Beijing 100871, China Kavli Institute for Astronomy and Astrophysics, Peking University, Beijing 100871, China    Yong Gao Department of Astronomy, School of Physics, Peking University, Beijing 100871, China Kavli Institute for Astronomy and Astrophysics, Peking University, Beijing 100871, China    Lijing Shao [email protected] Kavli Institute for Astronomy and Astrophysics, Peking University, Beijing 100871, China National Astronomical Observatories, Chinese Academy of Sciences, Beijing 100012, China    Ren-Xin Xu [email protected] Department of Astronomy, School of Physics, Peking University, Beijing 100871, China Kavli Institute for Astronomy and Astrophysics, Peking University, Beijing 100871, China
Abstract

The equation of state (EOS) of nuclear dense matter plays a crucial role in many astrophysical phenomena associated with neutron stars (NSs). Fluid oscillations are one of the most fundamental properties therein. NSs support a family of gravity gg-modes, which are related to buoyancy. We study the gravity gg-modes caused by composition gradient and density discontinuity in the framework of pseudo-Newtonian gravity. The mode frequencies are calculated in detail and compared with Newtonian and general-relativistic (GR) solutions. We find that the gg-mode frequencies in one of the pseudo-Newtonian treatments can approximate remarkably well the GR solutions, with relative errors in the order of 1%1\%. Our findings suggest that, with much less computational cost, pseudo-Newtonian gravity can be utilized to accurately analyze oscillation of NSs constructed from an EOS with a first-order phase transition between nuclear and quark matter, as well as to provide an excellent approximation of GR effects in core-collapse supernova (CCSN) simulations.

I Introduction

The oscillation modes of neutron stars (NSs) provide a means to probe the internal composition and state of dense matter. NSs have rich oscillation spectra, with modes associated with different physical origins, such as the internal ingredients, the elasticity of the crust, superfluid components, and so on [1]. For typical non-rotating fluid stars, the oscillation modes include the fundamental (ff), pressure (pp), and gravity (gg) modes, which provided the basic classification of modes according to the physics dominating their behaviours [2]. More realistic stellar models and rotation introduce additional classes of oscillation modes.

In this work, we study the gg-mode oscillations for non-rotating NSs in the framework of pseudo-Newtonian gravity [3, 4, 5, 6, 7, 8, 9, 10]. Reisenegger and Goldreich [11] investigated the gg-mode induced by composition (proton-to-neutron ratio) gradient in the cores of NSs. Moreover, hot young NSs may excite gg-modes supported by entropy gradients [12, 13, 14, 15]. It has also been demonstrated that the onset of superfluidity has a key influence on the buoyancy that supports the gg-modes [16, 17, 18, 19, 20]. Density discontinuity produced by abrupt composition transitions may play an important role in determining the gg-mode properties [21, 22]. Sotani et al. [23] calculated ff and gg modes of NSs with density discontinuity at an extremely high density and discussed the stability of the stellar models. A phase transition occurred in the cores of NSs with a polytropic equation of state (EOS) has been studied by Miniutti et al. [24]. The frequencies of gg-modes from density discontinuity are larger than those induced by the entropy gradient. Furthermore, discontinuity gg-mode may occur in perturbed quark-hadron hybrid stars [25, 26]. Recently, Zhao et al. [27] considered the gg-mode of NSs containing quark matter and discussed the Cowling approximation, which leads to a relative error of 10%\sim 10\% for higher-mass hybrid stars. We here focus on the ff and gg modes of NSs in pseudo-Newtonian gravity caused by the first-order phase transition in the cores of NSs.

The study of NS oscillations is timely in the gravitational-wave era [28, 29, 30]. Tidal interaction in a coalescing binary NS can resonantly excite the gg-mode oscillation of NSs when the frequency of the tidal driving force approaches the gg-mode frequencies [31, 32]. Moreover, the mixture of pure-inertial and inertial-gravity modes can become resonantly excited by tidal fields for rotating NSs [33, 34]. The gg-mode can also result in secular instability in rotating NSs [35]. Gaertig and Kokkotas [36] considered the gg-mode of fast-rotating stratified NSs using the relativistic Cowling approximation. The typical scenarios pertain to the pp-gg mode instability and the saturation of unstable modes [37, 38]. The universal relation of gg-mode asteroseismology has been discussed by Kuan et al. [39] for different classes of EOSs. In particular, the absence of very low-frequency gg-modes helps to explain the absence of tidal resonances [40]. The cut-off in the high-order gg-mode spectrum may also be relevant for scenarios of nonlinear mode coupling. The properties of gg-modes for newly-born strange quark stars and NSs using Cowling approximation in Newtonian gravity have been discussed by Fu et al. [41].

Hydrodynamical simulations are necessary to study the properties of the proto-NS in a core-collapse supernova (CCSN). The gg-mode of such a scenario may impact associated gravitational waves [42]. However, the physics of neutrino transport and EOS is very uncertain for the hydrodynamical simulations. As multi-dimensional general-relativistic (GR) codes for numerical simulations are scarce and have high demand of computational cost, most previous investigations relied on the Newtonian approximation for the strong gravitational field and fluid dynamics [3, 4]. Nevertheless, “Case A potential” formalism (c.f. Sec. II.1) was found to be a good approximation to relativistic solutions in simulating non-rotating or slowly rotating CCSNs. This potential allows for an accurate approximation of GR effects in an otherwise Newtonian hydrodynamic code, and it also works for cases of rapid rotation [4]. This has motivated a sequence of CCSN simulations [5, 6, 7, 8]. The effectiveness of using Case A potential formalism to approximate GR has been studied by Mueller et al. [4], Pajkos et al. [43], O’Connor et al. [7]. In particular, Mueller et al. [4] found that Case A potential formalism can not obtain the correct oscillation modes and indicated the failure of the Case A potential, possibly being attributed to the absence of a lapse function. Recently, Zha et al. [9] have extended the Case A potential formalism with a lapse function to simulate the oscillation of proto-neutron star (PNS). They found that Case A potential formalism with an additional lapse function can approximate well the frequency of the fundamental radial mode.

Tang and Lin [10] studied the radial and non-radial oscillation modes of NSs in pseudo-Newtonian gravity, including the Case A potential with and without the lapse function. Motivated by Tang and Lin [10], we here study the gg-mode of NS cores using Case A potential formalism with and without the lapse function. Our findings suggest that, with much less computational cost, pseudo-Newtonian gravity can be utilized to accurately analyze oscillation of NSs constructed from an EOS with a first-order phase transition, thus to provide an excellent approximation of GR effects in CCSN simulations.

The paper is organized as follows. In Sec. II, we introduce the key ingredients of the model, including different pseudo-Newtonian schemes and the buoyancy nature associated with gg-mode. The local dynamics of NS cores, including composition gradient and density discontinuity, are presented in Sec. III. Finally, we summarize our work in Sec. IV. Throughout the paper, we adopt geometric units with c=G=1c=G=1, where cc and GG are the speed of light and the gravitational constant, respectively.

II Key ingredients of the model

II.1 Case A potential in pseudo-Newtonian gravity

Case A effective potential is defined by replacing the Newtonian gravitational potential in a spherically symmetric Newtonian hydrodynamic simulation by [3, 10]

ΦTOV(r)=4πrdrr2(mTOV4π+r3P)×1Γ2(ρ+ρϱ+Pρ),\Phi_{\rm TOV}(r)=-4\pi\int^{\infty}_{r}\frac{\mathrm{d}r^{\prime}}{r^{\prime 2}}\left(\frac{m_{\rm TOV}}{4\pi}+r^{\prime 3}P\right)\times\frac{1}{\Gamma^{2}}\left(\frac{\rho+\rho\varrho+P}{\rho}\right)\,, (1)

where rr is the radial coordinate, ρ\rho is the rest-mass density, PP is the pressure, ϱ\varrho is the specific internal energy, and the total energy density is given by ϵ=ρ+ρϱ\epsilon=\rho+\rho\varrho. The function mTOVm_{\text{TOV}} is defined by

mTOV(r)=4π0rdrr2ϵΓ,m_{\rm TOV}(r)=4\pi\int^{r}_{0}\mathrm{d}r^{\prime}r^{\prime 2}\epsilon\Gamma\,, (2)

with

Γ=12mTOVr.\Gamma=\sqrt{1-2\frac{m_{\rm TOV}}{r}}\,. (3)

From Eq. (1) and Eq. (2), we have

dmTOVdr\displaystyle\frac{{\rm d}m_{\rm TOV}}{{\rm d}r} =4πr2ϵΓ,\displaystyle=4\pi r^{2}\epsilon\Gamma\,, (4)
dΦTOVdr\displaystyle\frac{{\rm d}\Phi_{\rm TOV}}{{\rm d}r} =4πr2(mTOV4π+r3P)1Γ2(ϵ+P)ρ.\displaystyle=\frac{4\pi}{r^{2}}\left(\frac{m_{\rm TOV}}{4\pi}+r^{3}P\right)\frac{1}{\Gamma^{2}}\frac{(\epsilon+P)}{\rho}\,. (5)

We use the Case A and Case A+lapse schemes and the other four schemes to study the gg-mode originating from the composition gradient and density discontinuity of NS cores in the framework of pseudo-Newtonian gravity. All background and perturbation equations for each scheme are given in the next three subsections and summarized in Table 1.

Table 1: Different schemes to calculate the oscillation modes, along with the corresponding background and the lapse function. Non-radial perturbation equations are the same [Eqs. (2831)] for all six schemes, but some of them include a lapse-function α\alpha in the hydrodynamic equations. Note that the lapse function only appears in the perturbation equations but not in the background equations.
Scheme Background equations Lapse function α\alpha
N Eqs. (6) to (8)
N+lapse Eqs. (6) to (8) Eq. (17)
TOV Eqs. (9) to (11)
TOV+lapse Eqs. (9) to (11) Eq. (17)
Case A Eqs. (12) to (14)
Case A+lapse Eqs. (12) to (14) Eq. (17)

II.2 Equilibrium configurations

We consider the following three sets of equilibrium configurations.

  1. (I)

    For the Newtonian (N) and Newtonian+lapse function (N+lapse) schemes, the hydrostatic equilibrium equations are

    dmdr\displaystyle\frac{{\rm d}m}{{\rm d}r} =4πr2ρ,\displaystyle=4\pi r^{2}\rho\,, (6)
    dPdr\displaystyle\frac{{\rm d}P}{{\rm d}r} =ρmr2,\displaystyle=-\frac{\rho m}{r^{2}}\,, (7)
    dΦdr\displaystyle\frac{{\rm d}\Phi}{{\rm d}r} =1ρdPdr.\displaystyle=-\frac{1}{\rho}\frac{{\rm d}P}{{\rm d}r}\,. (8)

    where ρ\rho is the rest-mass density, and Φ\Phi is the Newtonian gravitational potential.

  2. (II)

    Instead, if we consider spherical and static stars in GR, we have the Tolman-Oppenheimer-Volkoff (TOV) equations

    dmdr\displaystyle\frac{{\rm d}m}{{\rm d}r} =4πr2ϵ,\displaystyle=4\pi r^{2}\epsilon\,, (9)
    dPdr\displaystyle\frac{{\rm d}P}{{\rm d}r} =(ϵ+P)(m+4πr3P)r(r2m),\displaystyle=-\frac{(\epsilon+P)(m+4\pi r^{3}P)}{r(r-2m)}\,, (10)
    dΦdr\displaystyle\frac{{\rm d}\Phi}{{\rm d}r} =1ϵ+PdPdr.\displaystyle=-\frac{1}{\epsilon+P}\frac{{\rm d}P}{{\rm d}r}\,. (11)
  3. (III)

    Lastly, for the Case A and Case A+lapse schemes, the background equations are obtained by replacing the Newtonian gravitational potential by the Case A potential [3, 10], and we have

    dmdr\displaystyle\frac{{\rm d}m}{{\rm d}r} =4πr2ϵΓ,\displaystyle=4\pi r^{2}\epsilon\Gamma\,, (12)
    dPdr\displaystyle\frac{{\rm d}P}{{\rm d}r} =4πr2(m4π+r3P)1Γ2(ϵ+P),\displaystyle=-\frac{4\pi}{r^{2}}\left(\frac{m}{4\pi}+r^{3}P\right)\frac{1}{\Gamma^{2}}(\epsilon+P)\,, (13)
    dΦdr\displaystyle\frac{{\rm d}\Phi}{{\rm d}r} =1ρdPdr.\displaystyle=-\frac{1}{\rho}\frac{\mathrm{d}P}{\mathrm{d}r}\,. (14)

We use the modified Newtonian hydrodynamic equations in Zha et al. [9] and Tang and Lin [10], where a lapse function α\alpha is added to mimic the time-dilation effect. The modified hydrodynamic equations are:

ρt+(αρv)=0,\displaystyle\frac{\partial\rho}{\partial t}+\nabla\cdot(\alpha\rho\vec{v})=0, (15)
t(ρv)+[α(ρvv+PI)]=α(ρP)Φ,\displaystyle\frac{\partial}{\partial t}(\rho\vec{v})+\nabla\cdot[\alpha(\rho\vec{v}\vec{v}+P\overset{\leftrightarrow}{I})]=-\alpha(\rho-P)\nabla\Phi\,, (16)

where v\vec{v} is the fluid velocity and the lapse function is defined by

α=exp(Φ).\alpha=\exp(\Phi)\,. (17)

The readers can infer Tang and Lin [10] for a detailed variational derivation of the linearized fluid equations. We will use the same lapse function in our calculations.

II.3 Buoyancy and the gg-mode

As well known that NSs always have real frequency ff-mode and pp-mode regimes. However, gg-mode may have a real, imaginary, and zero frequency, which correspond to convective stability, instability, and marginal stability. We consider the local dynamics of NS cores, focusing on the buoyancy experienced by fluid elements and the associated gg-mode. The frequencies of gg-modes are closely related to the Brunt-Väisälä frequency NN, defined via

N2=gN2(1ce21cs2),N^{2}=g_{N}^{2}\left(\frac{1}{c_{\mathrm{e}}^{2}}-\frac{1}{c_{\mathrm{s}}^{2}}\right)\,, (18)

where gNg_{N} is the positive Newtonian gravitational acceleration, csc_{\mathrm{s}} is the adiabatic sound speed,

cs2=(Pρ)s.c_{\mathrm{s}}^{2}=\left(\frac{\partial P}{\partial\rho}\right)_{\mathrm{s}}\,. (19)

Here the subscript “s” means “adiabatic”, which in this case implies constant composition. The quantity cec_{\mathrm{e}} is given by

ce2=dPdρ,c_{\mathrm{e}}^{2}=\frac{{\rm d}P}{{\rm d}\rho}\,, (20)

where the subscript “e” stands for “equilibrium”. If cs2=ce2c_{\mathrm{s}}^{2}=c_{\mathrm{e}}^{2}, the star exhibits no convective phenomena (zero-buoyancy case). In this work, we consider only the gg-mode of NS cores, so we set cs2=ce2c_{\mathrm{s}}^{2}=c_{\mathrm{e}}^{2} for the crustal region. Again, cs2>ce2c_{\mathrm{s}}^{2}>c_{\mathrm{e}}^{2} (cs2<ce2c_{\mathrm{s}}^{2}<c_{\mathrm{e}}^{2}) denotes convective stability (instability). Combining Eqs. (1820), we can write the Brunt-Väisälä frequency as

N2=AgN,N^{2}=-Ag_{N}\,, (21)

where AA is

A=dlnρdr1Γ1dlnPdr,A=\frac{{\rm d}\ln\rho}{{\rm d}r}-\frac{1}{\Gamma_{1}}\frac{{\rm d}\ln P}{{\rm d}r}\,, (22)

which is called the Schwarzschild discriminant. If the star model obeys a simple polytropic EOS, P=KργP=K\rho^{\gamma}, then γ=dlnP/dlnρ\gamma=\rm d\ln P/\rm d\ln\rho is defined for the unperturbed background configuration. Hence, the Schwarzschild discriminant becomes

A=(1γ1Γ1)dlnPdr.A=\left(\frac{1}{\gamma}-\frac{1}{\Gamma_{1}}\right)\frac{{\rm d}\ln P}{{\rm d}r}\,. (23)

Clearly, if the adiabatic index Γ1>γ\Gamma_{1}>\gamma, the star is related to the convective stability, in the case of cs2>ce2c_{\mathrm{s}}^{2}>c_{\mathrm{e}}^{2}. In Sec. III.1, we will calculate the frequencies of gg-modes for the composition gradient, which is related to the discussion here.

II.4 Non-radial perturbation equations

In this section, we study non-radial oscillations of NSs in pseudo-Newtonian gravity. Tang and Lin [10] calculated the quadrupole (=2\ell=2) ff and pp modes. The perturbation of scalars is expanded in spherical harmonics and the Lagrangian displacement is expanded in vector spherical harmonics [13, 10]. When considering an eigenmode, we have

δρ=δρ~(r)Ym,\displaystyle\delta\rho=\delta\tilde{\rho}(r)Y_{\ell m}\,, (24)
δP=δP~(r)Ym,\displaystyle\delta P=\delta\tilde{P}(r)Y_{\ell m}\,, (25)
δΦ=δΦ~(r)Ym,\displaystyle\delta\Phi=\delta\tilde{\Phi}(r)Y_{\ell m}\,, (26)
ξ=U(r)Ymr^+V(r)Ym,\displaystyle\vec{\xi}=U(r)Y_{\ell m}{\hat{r}}+V(r)\nabla Y_{\ell m}\,, (27)

where YmY_{\ell m} is the standard spherical harmonic function, and r^\hat{r} is the radial unit vector. Then one can obtain the following system of equations for the fluid perturbations (see Tang and Lin [10], for a detailed variational derivation),

dUdr\displaystyle\frac{{\rm d}U}{{\rm d}r} =(2r+dΦdr+1γPdPdrAα)U+[α(+1)ρr2ω21αΓ1P]δP~\displaystyle=-\left(\frac{2}{r}+\frac{{\rm d}\Phi}{{\rm d}r}+\frac{1}{\gamma P}\frac{{\rm d}P}{{\rm d}r}-\frac{A}{\alpha}\right)U+\left[\frac{\alpha\ell(\ell+1)}{\rho r^{2}\omega^{2}}-\frac{1}{\alpha\Gamma_{1}P}\right]\delta\tilde{P}
+α(+1)r2ω2δΦ~,\displaystyle\quad+\frac{\alpha\ell(\ell+1)}{r^{2}\omega^{2}}\delta\tilde{\Phi}\,, (28)
dδP~dr\displaystyle\frac{{\rm d}\delta\tilde{P}}{{\rm d}r} =(ρω2αdPdrA)U+1Γ1PdPdrδP~ρdδΦ~dr,\displaystyle=\left(\frac{\rho\omega^{2}}{\alpha}-\frac{{\rm d}P}{{\rm d}r}A\right)U+\frac{1}{\Gamma_{1}P}\frac{{\rm d}P}{{\rm d}r}\delta\tilde{P}-\rho\frac{{\rm d}\delta\tilde{\Phi}}{{\rm d}r}\,, (29)
dδΦ~dr\displaystyle\frac{{\rm d}\delta\tilde{\Phi}}{{\rm d}r} =Ψ,\displaystyle=\Psi\,, (30)
dΨdr\displaystyle\frac{{\rm d}\Psi}{{\rm d}r} =2rΨ+(+1)r2δΦ~+4πρΓ1PδP~4πρAU.\displaystyle=-\frac{2}{r}\Psi+\frac{\ell(\ell+1)}{r^{2}}\delta\tilde{\Phi}+4\pi\frac{\rho}{\Gamma_{1}P}\delta\tilde{P}-4\pi\rho AU\,. (31)

To solve these equations, we require the boundary conditions at the center and surface of the NS. At the center, the regularity conditions of the variables yield the following relations [44, 10]

U=r1A0,\displaystyle U=r^{\ell-1}A_{0}\,, (32)
δP~=rB0,\displaystyle\delta\tilde{P}=r^{\ell}B_{0}\,, (33)
δΦ~=rC0,\displaystyle\delta\tilde{\Phi}=r^{\ell}C_{0}\,, (34)
Ψ=r1C0,\displaystyle\Psi=\ell r^{\ell-1}C_{0}\,, (35)
A0=αρω2(B0+ρC0),\displaystyle A_{0}=\frac{\alpha\ell}{\rho\omega^{2}}(B_{0}+\rho C_{0})\,, (36)

where B0B_{0} and C0C_{0} are constants. At the surface of the star, the perturbed pressure must vanish, which provides

dPdrU+δP~=0.\frac{{\rm d}P}{{\rm d}r}U+\delta\tilde{P}=0\,. (37)

The δΦ~\delta\tilde{\Phi} and dδΦ~/dr{{\rm d}\delta\tilde{\Phi}}/{{\rm d}r} are continuous, so we obtain

Ψ=+1rδΦ~.\Psi=-\frac{\ell+1}{r}\delta\tilde{\Phi}\,. (38)

Note that in the N, Case A, and TOV schemes, the lapse function equals to 1 (α=1\alpha=1).

Table 2: Comparison of the non-radial mode frequencies (unit: Hz) of a polytropic star model where polytropic index γ=2\gamma=2, K=1.4553×105g1cm5s2K=1.4553\times 10^{5}\ \rm g^{-1}\,cm^{5}\,s^{-2}, and central density ρc=7.9×1014gcm3\rho_{c}=7.9\times 10^{14}\ \rm g\,cm^{-3}, to earlier results of Westernacher-Schneider [44] and Tang and Lin [10].
Mode Westernacher-Schneider [44] Tang and Lin [10] Γ1=2.01\Gamma_{1}=2.01 Γ1=2.05\Gamma_{1}=2.05 Γ1=2.1\Gamma_{1}=2.1 Γ1=2.15\Gamma_{1}=2.15
p2p_{2} 7290 7932 7957 8049 8163 8276
p1p_{1} 5122 5131 5151 5216 5297 5377
ff 2024 2021 2021 2025 2029 2032
g1g_{1} 143 317 441 532
g2g_{2} 99 219 306 369
g3g_{3} 76 169 235 284

To test our numerical code, we have redone calculations with the same polytropic EOS as that in the Appendix A of Marek et al. [3], where the polytropic index γ\gamma and the adiabatic index Γ1>γ\Gamma_{1}>\gamma are constant throughout the stellar interior. Detailed numerical results are shown in Table 2. It is noted that our numerical results for the polytropic model with Γ1=γ\Gamma_{1}=\gamma agree with Table 3 of Tang and Lin [10]. In Table 2, we compare the frequencies of pp, ff, and gg modes computed with Γ1>γ\Gamma_{1}>\gamma and Γ1=γ\Gamma_{1}=\gamma [44, 10]. The frequencies of pp and ff modes increase with the increase of the adiabatic index Γ1\Gamma_{1}. In particular, the gg-mode frequencies also increase with increase of the adiabatic index Γ1\Gamma_{1}, which indicates a larger buoyancy.

III NUMERICAL RESULTS

III.1 Composition gradient

Taking the matter composition into account, and assuming that the model accounts for the presence of neutrons, protons, and electrons, we have a two-parameter EOS, P=P(n,x)P=P(n,x), which is a function of the baryon number density nn and the proton fraction x=np/nx=n_{\mathrm{p}}/n. Specifically, we use shorthand notations: “n\mathrm{n}” for neutrons, “p\mathrm{p}” for protons, and “e\mathrm{e}” for electrons. The energy per baryon of the nuclear matter can be written as [45, 46, 47, 31]

En(n,x)=Tn(n,x)+V0(n)+V2(n)(12x)2,E_{n}(n,x)=T_{n}(n,x)+V_{0}(n)+V_{2}(n)(1-2x)^{2}\,, (39)

where

Tn(n,x)=3522mn(3π2n)2/3[x5/3+(1x)5/3],T_{n}(n,x)={\frac{3}{5}\frac{\hbar^{2}}{2m_{\mathrm{n}}}(3\pi^{2}n)^{2/3}[x^{5/3}}+(1-x)^{5/3}]\,, (40)

is the Fermi kinetic energy of the nucleons, and mnm_{n} is the nucleon mass. V0V_{0} mainly specifies the bulk compressibility of the matter, and V2V_{2} is related to the symmetry energy of nuclear matter [48].

To compare the results of gg-modes in Newtonian gravity [31], we adopt the same V0V_{0} and V2V_{2} for different EOS models, based on the microscopic calculations in Wiringa et al. [47]. Detailed numerical results of V0V_{0} and V2V_{2} have been tabulated in Table IV of Wiringa et al. [47]. The approximate formulae of V0V_{0} and V2V_{2} are presented in Sec. 4.3 of Lai [31].

In this work, we consider the model “AU” (the EOS based on nuclear potential AV14+UVII in Wiringa et al. [47]) and the model “UU” (the EOS based on nuclear potential UV14+UVII in Wiringa et al. [47]), respectively. For the model AU, V0V_{0} and V2V_{2} (in the unit of MeV) are fitted as [31]

V0\displaystyle V_{0} =43+330(n0.34)2,\displaystyle=-43+330\,(n-0.34)^{2}\,, (41)
V2\displaystyle V_{2} =21n0.25,\displaystyle=21\,n^{0.25}\,, (42)

where nn is the baryon number density in fm3\rm fm^{-3}. For the model UU, we have

V0\displaystyle V_{0} =40+400(n0.3)2,\displaystyle=-40+400\,(n-0.3)^{2}\,, (43)
V2\displaystyle V_{2} =42n0.55.\displaystyle=42\,n^{0.55}\,. (44)

These fitting formulae are valid for 0.07fm3n1fm30.07\,{\rm fm}^{-3}\leq n\leq 1\,\rm fm^{-3}. For densities 0.001fm3<n<0.07fm30.001\,{\rm fm}^{-3}<n<0.07\,\rm fm^{-3}, we employ the EOS of Baym et al. [49], while for n0.001fm3n\leq 0.001\,\rm fm^{-3}, we employ the EOS of Baym et al. [50].

Once we have this relation, we can work out the mass-energy density, pressure, and adiabatic sound speed. The equilibrium configuration must satisfy the beta equilibrium,

μn=μp+μe,\mu_{\mathrm{n}}=\mu_{\mathrm{p}}+\mu_{\mathrm{e}}\,, (45)

and the charge neutrality

np=ne,n_{\mathrm{p}}=n_{\mathrm{e}}\,, (46)

where μi\mu_{i} are the chemical potentials of the three species of particles. The equilibrium proton fraction x(n)=xe(n)x(n)=x_{\mathrm{e}}(n) can be obtained by solving Eqs. (4.12–4.14) of Lai [31]. Hence, the mass-energy density and pressure are determined as

ϵ(n,x)\displaystyle\epsilon(n,x) =n[mn+E(n,x)/c2],\displaystyle=n\big{[}m_{n}+E(n,x)/c^{2}\big{]}\,, (47)
P(n,x)\displaystyle P(n,x) =n2E(n,x)n=2n3Tn+n3Te+n2[V0+V2(12x)2],\displaystyle=n^{2}{\frac{\partial E(n,x)}{\partial n}}=\frac{2n}{3}T_{n}+\frac{n}{3}T_{\mathrm{e}}+n^{2}\left[V_{0}^{\prime}+V_{2}^{\prime}\,(1-2x)^{2}\right]\,, (48)

where

Te(n,xe)=34c(3π2n)1/3xe4/3,T_{\mathrm{e}}(n,x_{\mathrm{e}})=\frac{3}{4}\hbar c(3\pi^{2}n)^{1/3}x_{\mathrm{e}}^{4/3}\,, (49)

is the energy per baryon of relativistic electrons. Here, and in the following, primes denote baryon number density nn derivatives (for example, V0=dV0/dnV_{0}^{\prime}={\rm d}V_{0}/{{\rm d}n}). The adiabatic sound speed cs2c_{\mathrm{s}}^{2} is

cs2=Pϵ=nϵ+P/c2Pn=nϵ+P/c2{109Tn+49Te+2n[V0+V2(12x)2]}+nϵ+P/c2{n2[V0′′+V2′′(12x)2]}.c_{\mathrm{s}}^{2}=\frac{\partial P}{\partial\epsilon}=\frac{n}{\epsilon+P/c^{2}}\frac{\partial P}{\partial n}\\ \quad=\frac{n}{\epsilon+P/c^{2}}\left\{{\frac{10}{9}}T_{n}+{\frac{4}{9}}T_{\mathrm{e}}+2n\left[V_{0}^{\prime}+V_{2}^{\prime}\,(1-2x)^{2}\right]\right\}\\ +\frac{n}{\epsilon+P/c^{2}}\left\{n^{2}\left[V_{0}^{\prime\prime}+V_{2}^{\prime\prime}\,(1-2x)^{2}\right]\right\}\,. (50)

The difference between cs2c_{\mathrm{s}}^{2} and ce2c_{\mathrm{e}}^{2} is given by

cs2ce2=nϵ+P/c2(PndPdn)=nϵ+P/c2(Px)dxdn=n3ϵ+P/c2[n(μe+μpμn)]dxdn.c_{\mathrm{s}}^{2}-c_{\mathrm{e}}^{2}=\frac{n}{\epsilon+P/c^{2}}\left(\frac{\partial P}{\partial n}-\frac{{\rm d}P}{{\rm d}n}\right)=-\frac{n}{\epsilon+P/c^{2}}\left(\frac{\partial P}{\partial x}\right){\frac{{\rm d}x}{{\rm d}n}}\\ =-\frac{n^{3}}{\epsilon+P/c^{2}}\left[{\frac{\partial}{\partial n}}(\mu_{\mathrm{e}}+\mu_{\mathrm{p}}-\mu_{\mathrm{n}})\right]{\frac{{\rm d}x}{{\rm d}n}}\,. (51)

From the beta equilibrium [i.e. Eq. (45)], we obtain

dxdn=[n(μe+μpμn)][x(μe+μpμn)]1.{\frac{{\rm d}x}{{\rm d}n}}=-\left[{\frac{\partial}{\partial n}}(\mu_{\mathrm{e}}+\mu_{\mathrm{p}}-\mu_{\mathrm{n}})\right]\left[{\frac{\partial}{\partial x}}(\mu_{\mathrm{e}}+\mu_{\mathrm{p}}-\mu_{\mathrm{n}})\right]^{-1}\,. (52)

Finally, the difference between cs2c_{\mathrm{s}}^{2} and ce2c_{\mathrm{e}}^{2} can be represented as

cs2ce2=n3ϵ+P/c2[n(μe+μpμn)]2[x(μe+μpμn)]1.c_{\mathrm{s}}^{2}-c_{\mathrm{e}}^{2}=\frac{n^{3}}{\epsilon+P/c^{2}}\left[{\frac{\partial}{\partial n}}(\mu_{\mathrm{e}}+\mu_{\mathrm{p}}-\mu_{\mathrm{n}})\right]^{2}\left[{\frac{\partial}{\partial x}}(\mu_{\mathrm{e}}+\mu_{\mathrm{p}}-\mu_{\mathrm{n}})\right]^{-1}\,. (53)
Refer to caption
Figure 1: The left panels show the pressure PP (upper) and the proton fraction x=np/nx=n_{\mathrm{p}}/n (lower) versus the mass-energy density ϵ\epsilon for representative EOS models AU and UU. The right panels show the relation between the adiabatic sound speed csc_{\mathrm{s}} and the fractional difference between cs2c_{\mathrm{s}}^{2} and ce2c_{\mathrm{e}}^{2} versus the mass-energy density ϵ\epsilon. The purple dashed line is the mass-energy density ϵ=0.07fm3\epsilon=0.07\,\rm fm^{-3}.

In the upper left panel of Fig. 1, we show the EOS models AU and UU, which include below neutron-drip region [49] and the lower-density crustal region [50]. In the bottom left panel of Fig. 1, we show the relation between the proton fraction x=np/nx=n_{\mathrm{p}}/n and the mass-energy density ϵ\epsilon. One notices that the value of xx of model UU is larger than that of model AU. In the right panels of Fig. 1, we show the relation between the adiabatic sound speed csc_{\mathrm{s}} and the fractional difference between cs2c_{\mathrm{s}}^{2} and ce2c_{\mathrm{e}}^{2}, as functions of the mass-energy density. Note that, in our work, we consider only gg-mode of the NS core, so we set cs2=ce2c_{\mathrm{s}}^{2}=c_{\mathrm{e}}^{2} in the lower-density region. As mentioned in Sec. 4 of Lai [31] that cs2=ce2c_{\mathrm{s}}^{2}=c_{\mathrm{e}}^{2} in the crustal region indicates effectively suppressing the crustal gg-mode while concentrating on the core gg-mode.

Refer to caption
Figure 2: Mass and radius of model UU as a function of central density ϵc\epsilon_{c}. The Case A and GR lines represent the background equations calculated by the Case A and TOV schemes, respectively.

As shown in Fig. 2, the mass and radius of model UU are plotted against the central density ϵc\epsilon_{c}. The Case A and GR lines represent the background equations calculated by the Case A and TOV schemes, respectively (see Table 1). We can see that the masses computed in the Case A formulation can approximate well the GR solutions. The Case A formulation has absolute percentage differences 5517%17\% for the radius of model UU. Note that the percentage difference of the stellar radii depends on the value of central density and the different EOS models. Detailed percentage differences of the stellar radii are illustrated in Appendix B of Tang and Lin [10].

The energy density profiles of the GR and Case A schemes with central density ϵ=2.0×1015gcm3\epsilon=2.0\times 10^{15}\rm g\,cm^{-3} for model UU is shown in Fig. 3. Compared with the GR solution, the Case A solution has a noticeable deviation only in the outer region of the surface. The total mass of the star is mainly determined by the high-density inner region. The above results may explain the fact that the total mass computed in the Case A formulation approximates well the GR solutions, though the radius has a large deviation.

Refer to caption
Figure 3: Comparison of the energy density profiles of the GR and Case A background solutions for model UU.

Note that the rest-mass density ρ\rho appears in the background and perturbation equations in N and N+lapse schemes; the total energy density ϵ\epsilon and rest-mass density ρ\rho exhibit the background equations in Case A and Case A+lapse schemes, but the rest-mass density ρ\rho appears in the perturbation equations. To compare with the results of Lai [31], we use the energy density ϵ\epsilon to obtain the mass-radius relation, as well as to solve perturbation equations. The difference between Case A and GR is apparent, though much smaller than the difference between Newtonian gravity and GR. Case A potential has captured some main effects from the full GR. As we will see, the perturbation results will be even closer to that of GR than the background results.

Lai [31] investigated ff and gg mode frequencies of EOS models AU and UU with a given mass M=1.4MM=1.4\,M_{\odot}111Lai [31] also calculated models UT and UU2. However, the maximum mass of the model UT does not accord with the new observation results [51, 52]. Also the model UU2 only considers the free n,p,e\mathrm{n},\mathrm{p},\mathrm{e} (V0=V2=0V_{0}=V_{2}=0). We will not include the two EOSs in our calculations.. They found that the ff-mode properties are very similar, due to the fact that the two EOSs have similar bulk properties (V0V_{0}) for the nuclear matter. However, the properties of the gg-mode are very different from models AU and UU. From the bottom right panel of Fig. 1, we find that the value of (cs2ce2)/cs2(c_{\mathrm{s}}^{2}-c_{\mathrm{e}}^{2})/c_{\mathrm{s}}^{2} is different with increase of the energy density. These differences reflect the sensitive dependence of gg-mode on the nuclear matter’s symmetry energy (V2V_{2}).

Refer to caption
Figure 4: The gg-mode frequencies for the EOS model AU, with a given mass M=1.98MM=1.98\,M_{\odot}. The upper panel shows the frequencies of the first eight quadrupolar (=2\ell=2) gg-modes with different schemes. The lower panel shows the absolute fractional difference ΔC\Delta_{\rm C} between our numerical results and the Case A+lapse scheme.
Refer to caption
Figure 5: Same as Fig. 4, bur for the EOS model UU.

In our study, we extend calculations in Lai [31] by computing the gg-mode. We use the stars with a fixed mass M=1.98MM=1.98\,M_{\odot} as an example. In the upper panel of Fig. 4, we plot the frequencies of the first eight quadrupolar gg-mode for the EOS AU. The results computed by all perturbation schemes are represented by different color lines in Fig. 4. The lower panel of Fig. 4 shows the absolute fraction difference ΔC\Delta_{\rm C} defined by

ΔC=|ffCase A+lapsefCase A+lapse|,\Delta_{\rm C}=\left\lvert\frac{f-f_{\text{Case A+lapse}}}{f_{\text{Case A+lapse}}}\right\rvert\,, (54)

where ff is the frequency of gg-mode obtained by our perturbation schemes in Table 1. According to the numerical results of non-radial oscillation (ff-mode) in Tang and Lin [10], the Case A+lapse scheme can approximate to about a few percents for a given mass M=1.4MM=1.4\,M_{\odot} in full GR. Hence, we use the results of Case A+lapse as the baseline in the case of the composition gradient. In particular, we found that the TOV+lapse scheme can give a good approximation to the gg-mode frequencies to a few percent levels. Besides, the absolute percentage difference ΔC\Delta_{\rm C} of the TOV+lapse scheme decreases with increasing nodes. We also plot the results of frequencies of gg-mode and the absolute percentage difference ΔC\Delta_{\rm C} for the EOS UU in Fig. 5. We seen similar properties of gg-mode, as the EOS AU in Fig. 4.

III.2 Density discontinuity

In this subsection, we study the effect of discontinuities at high density on the oscillation spectrum of a NS. We consider a simple polytropic EOS of the form [21, 22, 24]

P={Kϵγ,ϵ>ϵd+Δϵ,K(1+Δϵϵd)γϵγ,ϵϵd,P=\left\{\begin{array}[]{rl}K\epsilon^{\gamma}\,,&\epsilon>\epsilon_{\rm{d}}+\Delta\epsilon\,,\\ \displaystyle{K\left(1+\frac{\Delta\epsilon}{\epsilon_{\rm{d}}}\right)^{\gamma}\epsilon^{\gamma}}\,,&\epsilon\leq\epsilon_{\rm{d}}\,,\end{array}\right. (55)

where the discontinuity of amplitude Δϵ\Delta\epsilon is located at a mass-energy density ϵd\epsilon_{\rm d}. We study the properties of gg-modes with density discontinuity using the pseudo-Newtonian gravity schemes in Table 1.

Refer to caption
Figure 6: EOSs with density discontinuity, for different values of Δϵ/ϵd\Delta\epsilon/\epsilon_{\rm{d}}. The density and pressure are normalized by the standard nuclear density ϵnuc=2.68×1014gcm3\epsilon_{\rm nuc}=2.68\times 10^{14}\ \rm g\ cm^{-3}.
Refer to caption
Figure 7: (Left) The relation between the mass of NSs and central density ϵc\epsilon_{c} for different values of Δϵ/ϵd\Delta\epsilon/\epsilon_{\rm{d}}. (Right) Mass and radius relation of NSs with the same value of Δϵ/ϵd\Delta\epsilon/\epsilon_{\rm{d}}. The horizontal blue line shows the mass M=1.4MM=1.4\,M_{\odot}.

Now we have five parameters for a NS: the central density ϵc\epsilon_{c}, the discontinuity of amplitude Δϵ\Delta\epsilon, the critical density ϵd\epsilon_{\rm d}, the polytropic index γ\gamma, and KK. To compared with the results of non-radial oscillating relativistic stars in the full theory [i.e. without the relativistic Cowling approximation, 24], we adopt the same parameters as Miniutti et al. [24]: the polytropic index γ=2\gamma=2, K=180km2K=180\ \rm km^{2} for the NSs without discontinuity, and K(1+Δϵ/ϵd)2=180km2K(1+\Delta\epsilon/\epsilon_{\rm{d}})^{2}=180\ \rm km^{2} for the case with a discontinuity. Some examples of this EOS are illustrated in Fig. 6.

Refer to caption
Figure 8: The top panels plot the frequency of the non-radial ff-mode for different values of Δϵ/ϵd\Delta\epsilon/\epsilon_{\rm{d}} versus the density ϵd\epsilon_{\rm d} for the four perturbation schemes. The bottom panels show the absolute fractional difference ΔD\Delta_{\rm D} between our numerical results and the results of Miniutti et al. [24]. We here consider stars with a fixed mass M=1.4MM=1.4\,M_{\odot}

.

Refer to caption
Figure 9: Same as Fig. 8, but for the gg-mode frequencies.

In performing the calculation, boundary conditions must be specified at the locations of the density discontinuities. Finn [21] analyzed the jump conditions of the perturbation variables with the Cowling approximation in Newtonian gravity. Since the density is discontinuous, the perturbation variables are discontinuous as well, and the differential equations (2831) require jump conditions in the discontinuity density, denoted as [ρ\rho]

[U]=0,\displaystyle[U]=0\,, (56)
[δP~]=gN[ρ]U,\displaystyle[\delta\tilde{P}]=g_{N}[\rho]U\,, (57)
[δΦ~]=4π[ρ]U,\displaystyle[\delta\tilde{\Phi}]=-4\pi[\rho]U\,, (58)
[Ψ]=0.\displaystyle[\Psi]=0\,. (59)

To compare with the results of Miniutti et al. [24], we use the energy density ϵ\epsilon to solve perturbation equations.

In the left panel of Fig. 7, we show the mass MM versus central density ϵc\epsilon_{c} for each value of Δϵ/ϵd\Delta\epsilon/\epsilon_{\rm{d}}. As Δϵ/ϵd\Delta\epsilon/\epsilon_{\rm{d}} gets larger, the maximum mass decreases, and the stable region dM/dϵc>0\mathrm{d}M/\rm d\epsilon_{c}>0 becomes narrower and moves to a high-density region. In this work, we study only stable NS models with dM/dϵc>0\rm dM/\rm d\epsilon_{c}>0. In our analysis, we fix the mass of a NS to M=1.4MM=1.4\,M_{\odot} as an example. In the right panel of Fig. 7, we plot the mass-radius relation for NSs with and without density discontinuity. In both cases, we set the polytropic index γ=2\gamma=2. Comparing to the same EOS for ϵ<ϵd\epsilon<\epsilon_{\rm{d}}, we adopt K=180km2K=180\ \rm km^{2} for the NS models without discontinuity, and K(1+Δϵ/ϵd)2=180km2K(1+\Delta\epsilon/\epsilon_{\rm{d}})^{2}=180\ \rm km^{2} for the NS models with discontinuity. We find that the maximum mass is lower for the model with a discontinuity. Because the softening of EOS affected by the discontinuity. NSs with a discontinuity are more compact than those without discontinuity for a fixed mass.

Now we will focus on the =2\ell=2 non-radial oscillation modes. In particular, we consider the quadrupolar fundamental ff-mode and gravity gg-mode. The frequency versus density ϵd\epsilon_{\rm{d}} for the fixed mass M=1.4MM=1.4\,M_{\odot} is shown in the top panel of Fig. 8. The results computed by the four different perturbation schemes are represented by different color lines in Fig. 8. The GR curves in the upper panel correspond to the results of full perturbation theory in GR [24]. Additionally, the absolute fraction difference ΔD\Delta_{\rm D} defined by

ΔD=|ffGRfGR|,\Delta_{\rm D}=\left\lvert\frac{f-f_{\text{GR}}}{f_{\text{GR}}}\right\rvert\,, (60)

is shown in the bottom panel of Fig. 8. The frequency of ff-mode of the Case A+lapse scheme decreases with increasing density ϵd\epsilon_{\rm d}, which is similar to the GR results in trend. Again, the Case A+lapse scheme is quite accurate for the frequency of the ff-mode. For the Δϵ/ϵd=0.3\Delta\epsilon/\epsilon_{\rm{d}}=0.3, the Case A+lapse scheme is not as good as that of the Δϵ/ϵd=0.1,0.2\Delta\epsilon/\epsilon_{\rm{d}}=0.1,0.2 cases, but it is still the best among the four perturbation schemes. Tang and Lin [10] calculated ff-mode using Newtonian, Newtonian+lapse, Case A, and Case A+lapse schemes. They found that the Case A+lapse scheme performs much better and can reasonably approximate the ff-mode frequency.

Refer to caption
Figure 10: The top panels plot the frequency of the non-radial ff-mode for different values of Δϵ/ϵd\Delta\epsilon/\epsilon_{\rm{d}} versus the density ϵd\epsilon_{\rm d} for the four perturbation schemes. The bottom panels show the absolute fractional difference ΔD\Delta_{\rm D} between our numerical results and the results of Sotani et al. [23]. We here consider stars with a fixed mass M=1.2MM=1.2\,M_{\odot}

.

Refer to caption
Figure 11: Same as Fig. 10, but for the gg-mode frequencies.
Refer to caption
Figure 12: The top panel plots the frequency of the non-radial ff-mode with Δϵ/ϵd=0.3\Delta\epsilon/\epsilon_{\rm{d}}=0.3 versus the density ϵd\epsilon_{\rm d} for the different perturbation schemes. The bottom panel shows the absolute fractional difference ΔD\Delta_{\rm D} between our numerical results and the results of Miniutti et al. [24]. Here we consider stars with a fixed mass M=1.4MM=1.4\,M_{\odot}.
Refer to caption
Figure 13: Same as Fig. 12, but for the gg-mode frequencies.

In the top panel of Fig. 9, we show the frequency of gg-mode as a function of the density ϵd\epsilon_{\rm d} for the four schemes and the results of Miniutti et al. [24]. We also plot the results of ΔD\Delta_{\rm D} for the four schemes at the bottom of Fig. 9. In particular, we find that the Case A+lapse scheme can approximate the gg-mode frequency of GR reasonably well [24]. The percentage difference ΔD\Delta_{\rm D} of gg-mode of the Case A+lapse scheme decreases with increasing Δϵ/ϵd\Delta\epsilon/\epsilon_{\rm{d}}. The Case A+lapse scheme provides the best approximation to the frequencies of ff and gg modes. For the same central density and discontinuity density, the radius of density discontinuity RdR_{\rm{d}} is larger than the radius RR of the Newtonian star. Hence, we ignore the N and N+lapse schemes of discontinuity gg-mode in this work. Numerical results of the different schemes are given in Tables 3 and 4.

To prove that the pseudo-Newtonian treatments can approximate well the GR solutions, we also calculate the ff and gg modes of mass M=1.2MM=1.2\,M_{\odot} for the different schemes. Detailed numerical results of the different schemes are given in Fig. 10, Fig. 11, as well as in Tables 5 and 6. We find that the pseudo-Newtonian gravity can accurately describe the oscillation of the relativistic NSs constructed from an EOS with a first-order phase transition.

For a given density ϵd\epsilon_{\rm d} and Δϵ/ϵd\Delta\epsilon/\epsilon_{\rm d}, we show our numerical results for the frequencies of ff and gg modes with four schemes and the GR scheme, where the GR results were calculated by Miniutti et al. [24], Sotani et al. [23]. They integrated the equations describing the polar, non-radial perturbations of a non-rotating star as formulated by Lindblom and Detweiler [53] and Detweiler and Lindblom [54].

Finally, Sotani et al. [23] used the Cowling approximation to calculate the ff and gg modes and compared them to the results obtained from full GR. The results computed by the different perturbation schemes are represented by different colored lines in Fig. 12 and Fig. 13. We can see that the Case A+lapse scheme provides the best approximation to the frequency of gg-mode when the central density increases.

Table 3: Comparison between the frequencies of ff-mode (unit: Hz) of Miniutti et al. [24] and the different schemes in Table 1, with a given mass M=1.4MM=1.4\,M_{\odot} and Γ=2\Gamma=2 with different central densities. The polytropic coefficient KK is K(1+Δϵ/ϵ)2=180km2K(1+\Delta\epsilon/\epsilon)^{2}=180\,\rm km^{2}.
ϵd(gcm3)\epsilon_{\rm d}\ (\rm g\,cm^{-3}) Δϵ/ϵd\Delta\epsilon/\epsilon_{\rm d} Miniutti et al. [24] Case A Case A+lapse TOV TOV+lapse
0.0 1666 2144 1673 2423 1863
3×10143\times 10^{14} 0.1 1998 2629 1984 3058 2257
4×10144\times 10^{14} 0.1 1962 2562 1942 2987 2213
5×10145\times 10^{14} 0.1 1915 2482 1892 2890 2155
6×10146\times 10^{14} 0.1 1857 2404 1842 2782 2089
7×10147\times 10^{14} 0.1 1792 2302 1777 2644 2004
8×10148\times 10^{14} 0.1 1723 2215 1720 2526 1929
9×10149\times 10^{14} 0.1 1670 2152 1678 2431 1864
4×10144\times 10^{14} 0.2 2408 3269 2359 3968 2765
5×10145\times 10^{14} 0.2 2330 3117 2273 3764 2658
6×10146\times 10^{14} 0.2 2226 2901 2149 3536 2532
7×10147\times 10^{14} 0.2 2088 2665 2006 3238 2362
8×10148\times 10^{14} 0.2 1901 2451 1871 2860 2137
9×10149\times 10^{14} 0.2 1680 2171 1692 2451 1881
5×10145\times 10^{14} 0.3 3216 4213 2859 6350 3829
6×10146\times 10^{14} 0.3 3039 3909 2708 5718 3585
7×10147\times 10^{14} 0.3 2831 3605 2547 5066 3305
8×10148\times 10^{14} 0.3 2553 3044 2236 4298 2938
9×10149\times 10^{14} 0.3 2002 2311 1783 3053 2254
Table 4: Same as Table 3, but for the gg-mode frequencies.
ϵd(gcm3)\epsilon_{\rm d}\ (\rm g\,cm^{-3}) Δϵ/ϵd\Delta\epsilon/\epsilon_{\rm d} Miniutti et al. [24] Case A Case A+lapse TOV TOV+lapse
0.0
3×10143\times 10^{14} 0.1 504 571 500 604 523
4×10144\times 10^{14} 0.1 567 660 570 695 596
5×10145\times 10^{14} 0.1 613 730 624 766 651
6×10146\times 10^{14} 0.1 644 786 665 820 690
7×10147\times 10^{14} 0.1 659 828 692 855 712
8×10148\times 10^{14} 0.1 658 858 708 874 720
9×10149\times 10^{14} 0.1 641 876 713 876 712
4×10144\times 10^{14} 0.2 840 987 834 1059 883
5×10145\times 10^{14} 0.2 912 1093 916 1168 969
6×10146\times 10^{14} 0.2 961 1173 976 1252 1032
7×10147\times 10^{14} 0.2 987 1229 1016 1305 1070
8×10148\times 10^{14} 0.2 979 1262 1034 1311 1071
9×10149\times 10^{14} 0.2 906 1240 1009 1240 1010
5×10145\times 10^{14} 0.3 1211 1445 1174 1647 1286
6×10146\times 10^{14} 0.3 1281 1556 1257 1758 1375
7×10147\times 10^{14} 0.3 1326 1642 1319 1835 1439
8×10148\times 10^{14} 0.3 1339 1667 1341 1862 1467
9×10149\times 10^{14} 0.3 1251 1572 1274 1726 1383
Table 5: Comparison between the frequencies of ff-mode (unit: Hz) of Sotani et al. [23] and the different schemes, with a given mass M=1.2MM=1.2\,M_{\odot} for different central densities.
ϵd(gcm3)\epsilon_{\rm d}\ (\rm g\,cm^{-3}) Δϵ/ϵd\Delta\epsilon/\epsilon_{\rm d} Sotani et al. [23] Case A Case A+lapse TOV TOV+lapse
8×10148\times 10^{14} 0.1 2575 3514 2574 4148 2952
1.2×10151.2\times 10^{15} 0.1 2745 3792 2726 4542 3159
1.6×10151.6\times 10^{15} 0.1 2818 3913 2788 4721 3249
8×10148\times 10^{14} 0.2 2982 4212 2967 5112 3464
1.2×10151.2\times 10^{15} 0.2 3230 4643 3180 5776 3774
1.6×10151.6\times 10^{15} 0.2 3386 4914 3303 6231 3971
8×10148\times 10^{14} 0.3 3588 5426 3565 6898 4267
1.2×10151.2\times 10^{15} 0.3 4046 6440 3970 6647 4890
Table 6: Same as Table 5, but for the gg-mode frequencies.
ϵd(gcm3)\epsilon_{\rm d}\ (\rm g\,cm^{-3}) Δϵ/ϵd\Delta\epsilon/\epsilon_{\rm d} Sotani et al. [23] Case A Case A+lapse TOV TOV+lapse
8×10148\times 10^{14} 0.1 727 856 730 901 761
1.2×10151.2\times 10^{15} 0.1 871 1076 891 1134 931
1.6×10151.6\times 10^{15} 0.1 963 1247 1004 1311 1048
8×10148\times 10^{14} 0.2 1041 1234 1033 1318 1084
1.2×10151.2\times 10^{15} 0.2 1271 1570 1277 1688 1349
1.6×10151.6\times 10^{15} 0.2 1439 1847 1463 1994 1554
8×10148\times 10^{14} 0.3 1272 1558 1260 1692 1329
1.2×10151.2\times 10^{15} 0.3 1595 2021 1575 2228 1674

IV Conclusions

In light of new observations, oscillating modes of NSs are of particular interests to the physics and astrophysics communities in recent years. In this work, we have investigated the properties of the gravity gg-mode for NSs in the framework of pseudo-Newtonian gravity. Tang and Lin [10] have investigated barotropic oscillations (Γ1=γ\Gamma_{1}=\gamma and the Schwarzschild discriminant A=0A=0). We extended the work and have studied the gg-mode of NSs with the same polytropic EOS model. We find that, the gg-mode frequencies increase with increasing adiabatic index, which indicates that the buoyancy becomes much larger.

A deeper understanding of the oscillation of NSs, which could be associated with emitted gravitational waves, requires an analysis of both the state and composition of the NS matter. We considered the case of the composition gradient, and have extended calculations in Lai [31] to compute the gg-mode. The value of (cs2ce2)/cs2(c_{\mathrm{s}}^{2}-c_{\mathrm{e}}^{2})/c_{\mathrm{s}}^{2} is different when the energy density increases. In particular, these differences reflect the sensitive dependence of gg-mode on the nuclear matter’s symmetry energy [V2V_{2} in Eq. (39))]. Note that the tidal deformability of binary NSs appears to be related to the dominant oscillation frequency of the post-merger remnant [55]. The impact of thermal and rotational effects can provide simple arguments that help explain the result [56]. More recently, Andersson et al. [57] consider the dynamic tides of NSs to build the structure NSs in the framework of post-Newtonian gravity. We may expect using the pseudo-Newtonian gravity to study the resonant oscillations and tidal response in coalescing binary NSs in the future.

We considered a phase transition occurring in the inner core of NSs, which could be associated with a density discontinuity. Phase transition would produce a softening of EOSs, leading to more compact NSs. Using the different schemes, we have calculated the frequencies of ff and gg modes for the =2\ell=2 component. Compared to the results of GR [24, 23], the Case A+lapse scheme can approximate the ff-mode frequency very well. The absolute percentage difference ΔD\Delta_{\rm D} ranges from 0.010.01 to 0.10.1 percent. In particular, we find that the Case A+lapse scheme also can approximate the gg-mode frequency of GR reasonably well [24, 23]. The percentage difference ΔD\Delta_{\rm D} of gg-mode of the Case A+lapse scheme decreases with increasing Δϵ/ϵd\Delta\epsilon/\epsilon_{\rm{d}} in our model.

The existence of a possible hadron-quark phase transition in the central regions of NSs is associated with the appearance of gg-mode, which is extremely important as they could signal the presence of a pure quark matter core in the center of NSs [58]. Our findings suggest that the pseudo-Newtonian gravity, with much less computational efforts than the full GR, can accurately study the oscillation of the relativistic NSs constructed from an EOS with a first-order phase transition. Observations of gg-mode frequencies with density discontinuity may thus be interpreted as a possible hint of the first-order phase transition in the core of NSs. Lastly, our work also provides more confidence in using the pseudo-Newtonian gravity in the simulations of CCSNs, thus reducing the computational cost significantly.

McDermott et al. [13] investigated the non-radial oscillation of NSs using Cowling approximation and discussed the different damping mechanisms. Reisenegger and Goldreich [11] considered the gg-mode induced by composition (proton-to-neutron ratio) gradient in the cores of NSs and discussed damping mechanisms. They also estimated damping rates for the core gg-modes. Cutler et al. [59] assessed the accuracy of Cowling eigenfunctions and found that the relativistic Cowling approximation by McDermott et al. [12] accurately predicts frequencies and eigenfunctions. Chugunov and Gusakov [60] calculated the non-radial oscillations of superfluid non-rotating stars. An approximate decoupling of equations describing the oscillation modes of superfluid and normal fluid has been studied by Gusakov and Kantor [61]. Further, Gusakov et al. [62] developed an approximate method to determine the eigenfrequencies and eigenfunctions of an oscillating superfluid NS. In this work, we found that the gg-mode frequencies in one of the pseudo-Newtonian treatments can approximate remarkably well the GR solutions than the relativistic Cowling approximation. Hence, we may conjecture that the eigenfunction and dissipation of the pseudo-Newtonian treatments are also more accurate than the relativistic Cowling approximation. Based on the approximate method of Gusakov et al. [62], one can calculate the eigenfunctions and the different damping mechanisms in future studies.

Acknowledgements.
We thank the anonymous referee for helpful comments, and Zexin Hu and Yacheng Kang for the helpful discussions. This work was supported by the National SKA Program of China (2020SKA0120300, 2020SKA0120100), the National Natural Science Foundation of China (11975027, 11991053), the National Key R&D Program of China (2017YFA0402602), the Max Planck Partner Group Program funded by the Max Planck Society, and the High-Performance Computing Platform of Peking University.

References